Agrobiodiversity - Platform for Agrobiodiversity Research

advertisement
Agrobiodiversity: Its Value, Measurement, and Conservation in the Context
of Sustainable Agriculture
Brian Love1
Dean Spaner2
ABSTRACT. Conservation of agrobiodiversity is an important component of sustainable
agriculture and is important internationally. Ex-situ conservation in genebanks has been the
dominant strategy. Recently, in-situ conservation has been advocated as a complementary
strategy. This review 1) defines the context of agrobiodiversity conservation, 2) discusses its
value and measurement, 3) explores the advantages and disadvantages of ex-situ and in-situ
conservation approaches, and 4) outlines the importance of seed exchange and ethical concerns.
KEYWORDS. agrobiodiversity, conservation, in-situ, ex-situ, seed exchange, genebanks, onfarm conservation
1
Brian Love is a Ph.D. student in plant science at the University of Alberta in the Department of Agricultural, Food,
and Nutritional Sciences, 4-10 Ag/For Building Edmonton, Alberta, Canada, T6G 2P5, and is presently a researcher
for the Native Species Reforestation Project (PRORENA) in Panama. (brianl@ualberta.ca).
2
Dean Spaner is an Associate Professor at the University of Alberta in the Department of Agricultural, Food, and
Nutritional Science, 4-10 Ag/For Building Edmonton, Alberta, Canada, T6G 2P5. Corresponding Author,
(dean.spanner@ualberta.ca).
A Natural Sciences and Engineering Research Council Canada Graduate Student Scholarship, and the John G. Bene
Fellowship in Community Forestry offered by the International Development Research Centre supported B. Love
during the writing of this review.
1
Agrobiodiversity: Its Value, Measurement, and Conservation in the Context
of Sustainable Agriculture
ABSTRACT. Conservation of agrobiodiversity is an important component of sustainable
agriculture and is important internationally. Ex-situ conservation in genebanks has been the
dominant strategy. Recently, in-situ conservation has been advocated as a complementary
strategy. This review 1) defines the context of agrobiodiversity conservation, 2) discusses its
value and measurement, 3) explores the advantages and disadvantages of ex-situ and in-situ
conservation approaches, and 4) outlines the importance of seed exchange and ethical concerns.
KEYWORDS. agrobiodiversity, conservation, in-situ, ex-situ, seed exchange, genebanks, onfarm conservation
2
INTRODUCTION
Agrobiodiversity
Biodiversity (biological diversity) is the variability among living organisms and the
ecological complexes (e.g. ecosystems) they compose (UNCED, 1992). Agrobiodiversity refers
to the diversity of living organisms (plants, animals, bacteria, etc.) used in agriculture (Wood and
Lenne, 1999). This review is limited to plant diversity and will focus on food plants. Thus, the
terms ‘crop genetic resources’ and ‘agrobiodiversity’ will be used interchangeably.
Agrobiodiversity underpins the development of sustainable agriculture (Cleveland et al., 1994;
Ceccarelli et al., 1992). Globally, there are an estimated 250 000 to 500 000 plant species (FAO,
1998). Of these, only 1 500 have been used in agriculture (Wilkes, 1993). Currently, there are
120 nationally important crops and three food crops; wheat (Triticum spp.), rice (Oryza spp.) and
maize (Zea mays L.) provide over half of the food energy consumed by humans (FAO, 1998).
While the number of crop species is low, most agrobiodiversity exists within crops at the varietal
level (Brush, 2004). Crop plant relatives are also an important component of agrobiodiversity
(Hawkes, 1977).
Sustainability
There is no widely accepted definition of sustainable agriculture (Lewandowski et al.,
1999). This review treats sustainable agriculture relative to its ecological conceptualization and
defines it as the management and utilization of agroecosystems in a manner that does not
degrade resources beyond recuperation, and permits indefinite future use by maintaining
biological integrity and functionality. Agrobiodiversity constitutes the biological underpinning
of agriculture (Fowler and Hodgkin, 2004). Experiments in natural and microcosm
environments have linked biodiversity to increased productivity (Tilman et al., 1996), increased
3
stability (McNaughton, 1977), and increased ecosystem functioning (Naeem et al., 1994). Thus,
agrobiodiversity contributes to the sustainability of agricultural systems.
In-Situ and Ex-Situ Conservation
Conservation strategies may be either ex-situ or in-situ (Maxted et al., 1997; UNCED,
1992). Ex-situ strategies conserve diversity outside of natural habitats, while in-situ strategies
conserve diversity in the setting where it developed (UNCED, 1992).
Ex-situ conservation includes genebank storage (seed and field), in-vitro storage, pollen
storage, and DNA storage (Maxted et al., 1997). Seed genebanks are the most common storage
practice (FAO, 1998). Generally, ex-situ conservation for plant breeding involves the collection,
classification, evaluation and utilization of agrobiodiversity (Marshall and Brown, 1975).
In-situ conservation includes conservation in reserves, on farms, and in home gardens
(Maxted et al., 1997). In this review on-farm and homegarden conservation are considered to be
the same and referred to as on-farm conservation. Reserve strategies apply to forests and wild
crop relatives and will only be considered historically. Iltis (1974) suggested a reserve strategy
for food crops in which the genetic landscape would be frozen by isolating it spatially and
temporally. However, traditional communities are agriculturally dynamic (Louette, 1999), often
rendering such a strategy untenable.
On-farm conservation encourages farmers to continue selection and management of local
crop populations (Brush, 1999). On-farm conservation has focused on de facto conservation in
centers of origin (Qualset et al., 1997; Brush, 1991), but exceptionally valuable varieties are
found outside their centers of origin (Vavilov, 1951). Resource-poor farmers in marginal
4
environments often maintain large amounts of agrobiodiversity (Maxted et al., 2002; Wood and
Lenne, 1997; Bellon, 1996).
Theoretical Underpinnings of Conservation
Population biology and genetics provide theoretical frameworks for ex-situ conservation
and in-situ conservation in reserves (Brush, 2004). In contrast, the scientific basis for on-farm
conservation has not been well developed (Wood and Lenne, 1997; Maxted et al. 2002),
although Louette (1999) has advocated the use of metapopulation theory for understanding onfarm conservation. Brush (2004) recognized the potential of niche theory and metapopulation
theory as explanatory frameworks, but warned that their development for natural populations
undergoing natural selection may reduce their utility for crop populations that undergo both
natural and artificial selection.
Past Reviews
Review articles have addressed in-situ (Hammer, 2003; Brush, 1999; Wood and Lenne,
1997; Altieri and Merrick, 1987) and ex-situ (Qualset and Shands, 2005; Wright, 1997;
Goodman, 1990) conservation, the measurement of biodiversity (Mohammadi and Prasanna,
2003; Glaubitz and Moran, 2000; Brown, 1978; Marshall and Brown, 1975; Peet, 1974), its
value (Brush and Meng, 1998; Bellon, 1996; Wilson and Ehrlich, 1991), seed exchange (Louette
et al., 1997; Almekinders et al., 1994), ethics (Evenson, 1999; Brush, 1992), and the role of
agrobiodiversity in sustainability (Cleveland et al., 1994; Ceccarelli et al., 1992). In part, this
review aims to provide an integration of these topics.
5
Historical Context
In ancient times (2500 BC) the Sumerians, Egyptians, and Chinese were all engaged in
plant introduction from abroad (Ryerson, 1933). In the modern era, especially during colonial
times, crop species continued to be collected, with botanical gardens being the primary
repository for collections (Maxted et al., 1997; Brockway, 1979). Today the agriculture of most
countries is based on foreign plant introductions (Fowler and Hodgkin, 2004).
Linneaus (1707-1778) formalized the classification of living organisms with his work
Systema Naturae (Dickinson, 1967). Mendel (1822-1884) discovered the principles of heredity
(Bateson, 1913). Darwin called attention to diversity (variation within and among living
organisms) and its link to heredity and selection pressure, and discussed agricultural diversity
(Darwin, 1883; Darwin, 1860). De Candolle suggested crop wild types as a proxy for identifying
centers of domestication (de Candolle, 1914). Vavilov (1951) thereafter suggested that centers
of crop diversity, coupled with the presence of wild types, indicated centers of origin. At the
same time, he pioneered large-scale, long-term, international collecting missions in the 1920s
and 30s (Pistorius, 1997).
Early discussion of crop genetic resource conservation dates back to 1890 (Zeven, 1998).
Harlan and Martini (1936) and Frankel (1954) raised concerns early on about the loss of crop
genetic resources. While the Green Revolution has often been cited as eroding genetic resources
(Almekinders and de Boef, 1999; Tilman, 1998; Matson et al., 1997) these assertions may be
unfounded (Smale, 1997; Wood and Lenne, 1997). Harlan (1975) described the international
response to these concerns. Eventually, the International Board of Plant Genetic Resources1
(IBPGR) was established in 1974, conducting ~500 collection missions in its first decade of
1
The IBPGR eventually became the International Plant Genetic Resources Institute.
6
existence (Lawrence, 1984). Worldwide, there are currently ~6 million accessions2 (1-2 million
unique accessions due to duplication) in over 1300 germplasm collections (FAO, 1998). Further
collection is still needed for minor species and to fill gaps in a few major species (e.g. maize)
(FAO, 1998). The Convention on Biological Diversity (UNCED, 1992) and the International
Treaty on Plant Genetic Resources for Food and Agriculture (FAO, 2001) created a formal
international legal mandate for agrobiodiversity conservation.
These agreements have highlighted in-situ approaches. Reserve strategies such as forest
reserves (FAO, 1998) or reserves for wild crop relatives (Frankel et al., 1995) have traditionally
been employed for in-situ conservation (Maxted et al., 2002; Brown, 1999). For instance the
Sierra de Manantalán Biosphere Reserve in Mexico was created to protect perennial teosinte
(Zea diploperennis) and the Garo Hills Sanctuary in India was developed to protect oranges
(Citrus indica) (Meilleur and Hodgkin, 2004). On-farm conservation requiring farmer
participation is now, also, considered an important in-situ strategy (Maxted et al., 2002;
Cromwell and van Oosterhout, 2000; Brush, 1999). While the theoretical benefits of this onfarm conservation have been well developed (Brush, 1999; Maxted et al., 1997), there are only a
relatively small number of projects worldwide (see Jarvis et al., 2000; FAO, 1998; Bretting and
Duvick, 1997). In-situ and ex-situ strategies are deemed complementary (UNCED, 1992; FAO,
1998), but few conservation programs employ both approaches (Maxted et al. 1997).
VALUE OF AGROBIODIVERSITY
Biodiversity may be valued because of ethical obligation, economic benefits, or
preservation of essential ecosystem services (Wilson and Ehrlich, 1991). Plant breeders and
farmers approach the value of agrobiodiversity from different perspectives.
2
An accession is a sample of a crop species taken from a given geographical location at a given time.
7
Germplasm is used by plant breeders to adapt crops to heterogeneous and changing
environments (Bellon, 1996). Agrobiodiversity conservation is important due to the need to
broaden the genetic base of crop plants (Cooper et al., 2001), to prevent the loss of uniquely
adapted ecotypes (Vavilov, 1957), and to develop crops for local adaptation (Ceccarelli, 1996).
The Irish potato famine (Bursh, 2004) and the Southern leaf blight of maize epidemic (Harlan,
1972) are examples of the hazards of a narrow genetic base. The Green Revolution of the 1960s
produced high yielding crop varieties and increased recognition of the value of crop genetic
resources (Brush, 2004; Pistorius, 1997).
Pardey et al. (1998) proposed that agrobiodiversity has three types of value: 1) use value,
2) option value, and 3) existence value. Use value is associated with genetic resources’ current
effect on yield. Option value is associated with a future unknown use (e.g. resistance to new
disease). Existence value is associated with the satisfaction people derive from simply knowing
that diversity exists (Bellon, 1996). Agrobiodiversity conservation is a prerequisite for
developing sustainable agriculture because it enables breeders to address changing environments
(e.g. climate change).
The monetary value of modern plant varieties is difficult to estimate and includes the
value of agrobiodiversity, plant breeders’ work, and other research inputs (FAO, 1998). Annual
global markets for products (including agricultural products) derived from genetic resources are
estimated to be worth US $500 - $800 billion (ten Kate and Laird, 2000). The incorporation of a
rice landrace’s genes into an improved rice variety has been estimated to be worth US $50
million (Evenson and Gollin, 1997).
8
Farmers may plant varieties to prevent their loss (Bellon, 1996), but generally
conservation of agrobiodiversity for its own sake is not a farmer objective (Meng et al., 1998a).
Still, de facto conservation by farmers is commonplace (Brush, 1991). Farmers use intraspecific
(varietal) diversity to cope with uncertain and heterogeneous farming conditions (Bellon, 1996).
Just and Zilberman (1983) theoretically demonstrated that planting multiple varieties minimizes
risk while maximizing mean economic returns. Yield stability is arguably a benefit of varietal
diversity (Cleveland et al., 1994; Ceccarelli et al., 1992). However, reviews of the performance
of mixtures (Trenbath, 1974; Marshall and Brown, 1973) have found that while mixtures
frequently outperform their average component, seldom do they outperform their best
component. However, the best component may vary from year to year, making mixtures more
stable across time. Greater stability (Cleveland et al., 1994) and the flexibility to address
heterogenous and marginal farming conditions (Bellon, 1996) are direct contributions of
agrobiodiversity to sustainability.
Diversity in varietal maturity assists in the scheduling of labor inputs (Zimmerer, 1991).
Varietal diversity may be important due to ritual (specific varieties for religious ceremonies) and
prestige (varieties valued for their novelty) (Brush, 2004). Both interspecific (crop species)
(Fleuret and Fleuret, 1980) and intraspecific (Bellon, 1996) diversity are important for dietary
diversification. Bellon (1996) developed a framework for assessing the dynamics of farmer
retention of intraspecific diversity. In this framework, farmers are considered to address their
farming concerns (soil types, dietary, etc.) by choosing varieties that best address each concern
(Bellon, 1996). More information is needed to understand if retaining intraspecific diversity is
due to a lack of appropriate improved varieties or because it meets special farmer requirements
(Brown, 1999).
9
MEASUREMENT OF AGROBIODIVERSITY
Diversity can be measured in physically classified units (e.g. species, races, varieties), or
in terms of genes. Genetic diversity exists between and within a number of different levels. This
review employs definitions adapted from Cleveland et al. (1999):
Species: Is made up of a number of populations that have the potential to interbreed. In
many cases species are able to breed with closely related species (Ellstrand, 2001) making
this classification somewhat subjective.
Race: Is a group of related varieties that share a suite of traits that define that race.
Variety: Is a distinct subunit of a crop species. Farmer varieties are subunits of crop species
distinct enough to be named by farmers and include both local and exotic materials of
improved and unimproved nature.
Population: A group of individual plants of a particular variety managed under the same
regime (e.g. farmer X’s yellow maize). A seedlot (Louette, 1999) is the portion of the
population that is used in any given year to regenerate the population through planting.
Landrace Defined
Landrace is a popular term in crop genetic resources literature. Despite its initial use in
1890, few authors’ have attempted to define it (Zeven, 1998). Landraces are thought to be: 1)
adapted to local conditions (Brush, 1999; FAO, 1998; Cleveland et al., 1994); 2) highly diverse
(Brown, 1999; Brush, 1999; FAO, 1998; Qualset et al., 1997; Cleveland et al., 1994); 3) tolerant
10
of abiotic and biotic stress (Zeven, 1998; Qualset et al., 1997); and 4) the product of farmer
selection (Swanson and Goeschl, 2000; FAO, 1998; Cleveland et al., 1994; Vavilov, 1957).
Wood and Lenne (1997) question whether landraces are locally adapted and argue that
local adaptation is unlikely unless farmers select for it or farming is carried out in stress-prone
environments. Louette (1999) questions the landrace definition by demonstrating that seed
exchange and cross-pollination make the notion of “local varieties” difficult to define.
Regardless, the distinction between the products of formal breeding programs and those arising
in agricultural ecosystems is an important one. The term farmer variety as defined above shall
be used in this review when discussing agrobiodiversity and includes landraces (populations
improved by generations of farmer management and selection), past and current modern varieties
(populations improved by formal breeding programs), and creolized varieties (populations
resulting from crosses between landraces and modern varieties while under farmer management).
Quantification of Diversity
Community ecology has focused on diversity at the species level (Hanski and Simberloff,
1997). Some agrobiodiversity projects have chosen to quantify diversity at the species level (e.g.
Zarin et al. 1999), but most agrobiodiversity work is aimed at diversity at the varietal level (e.g.
Brush, 2004; Salick et al., 1997; Bellon, 1996; Cleveland et al., 1994). Crop breeding’s
foundation on genetic principles makes the genetic diversity contained within populations of
varieties important (Charcosset and Moreau, 2004). Techniques for quantifying diversity at the
species level are well developed, and are applicable/amenable to quantifying diversity at the
varietal level. Quantification of genetic diversity is a rapidly evolving field and involves more
complicated measures and techniques.
11
Knowledge about the diversity contained in crop species and their varieties assists
breeders in crop improvement (Mohammadi and Prasanna, 2003). Agrobiodiversity assessment
requires estimates of the geographic area farmers’ crop species and varieties cover, population
size, and inherent genetic diversity (Brown, 1999). Such assessment is needed to develop both
ex-situ (Marshall and Brown, 1975) and in-situ (Brown, 1999) conservation strategies.
Measures of Diversity
Species/varietal diversity can be measured in a number of different ways. In ecology,
species diversity in communities is the object of measurement (Whittaker, 1965). Communities
consist of many coexisting populations, and community boundaries can be difficult to define. In
agriculture, crop communities are defined by the boundaries of a farmer’s field, making
community identification less subjective. Community censuses are difficult and samples in the
form of quadrats may be taken instead (Whittaker, 1965).
Species richness, total number of species/varieties in a defined space at a point in time
(Hubbell, 2001), is the simplest measure of diversity. In agriculture, this would be the total
number of crop species sampled within a farmer’s field. Species evenness, how many
individuals belong to each species in a community, is also a component of diversity (Margalef,
1958). Abundance-diversity curves graphically portray diversity as both richness and evenness
(Zarin et al., 1999). Relative species abundance measures how rare or common species/varieties
are (Hubbell, 2001). Unfortunately, it can be difficult to identify individuals in plant
communities (Whittaker, 1965) and high planting densities can make counting of individuals
inefficient. Percent cover is an alternative measure that is rapid, repeatable, and
methodologically robust (McCune and Grace, 2002).
12
Indices such as the Shannon-Wiener’s or Simpson’s index combine richness and
evenness values to produce a numerical output (McCune and Grace, 2002). There are a
multitude of indices for measuring diversity (Peet, 1974). Hurlbert (1971) suggests that such
indices are meaningless because they have been defined in so many different ways, and he
demonstrates that there is non-concordance between indices. While species richness does not
include a measure of evenness, it is simple and easy to communicate (Purvis and Hector, 2000).
The above measures of species diversity can be applied to other taxonomic units, such as crop
races or varieties.
Diversity measures are spatially (sample unit) and temporally specific. Three levels of
diversity have been recognized: alpha, beta, and gamma (Whittaker, 1965). Alpha diversity is an
estimate of the average diversity contained within a sampling unit (community, quadrat) for
which diversity is being measured (e.g. crop species, varieties, etc.). Beta diversity is a measure
of compositional change between sample units (e.g. number of crop species or varieties that
change from farmer to farmer). Gamma diversity is a measure of the diversity contained in a
number of units belonging to a larger unit (e.g. total number of crop species or varieties grown in
an agroecological zone). Alpha and gamma diversity employ the measures outlined above, while
beta diversity requires different measures (McCune and Grace, 2002). Condit et al. (2002) used
reduction in species similarity across increasing distance as a measure of beta diversity. Vellend
(2001) has discussed the measurement of beta diversity in terms of species turnover along
environmental gradients.
While spatial diversity is the most common conceptualization of diversity, turnover of
species/varieties in time (temporal diversity) is important and can substitute for lack of spatial
diversity (Meng et al., 1998b). Brennan and Byerlee (1991) provided a measure for the rate of
13
varietal replacement, which calculates the average age of varieties since release, weighted by the
area they cover.
Genetic Diversity Parameters
A number of different genetic parameters can be used to describe genetic diversity.
Hamrick and Godt (1996) used percent polymorphic loci, mean number of alleles per
polymorphic locus and Hardy-Weinberg expected heterozygosity averaged across all loci.
Proportion of total diversity among populations is also a commonly used genetic parameter
(Yang and Yeh, 1992). Wright (1951) developed a number of measures called F-statistics to
quantify genetic differentiation between and within populations (Mohammadi and Prasanna,
2003).
There are three F-statistics (FIT, FST, and FIS), which are types of inbreeding coefficients
and measure inbreeding in individuals relative to the total population (FIT), inbreeding in
subpopulations relative to the total population (FST), and inbreeding in individuals relative to
their subpopulations (FIS), respectively (Hartl and Clark, 1989). Computation of these
parameters is complex, especially if there are unequal sample and population sizes and there is
disagreement regarding the computation and interpretation of these parameters (Weir and
Cockerham, 1984).
Nei’s GST is an alternative for FST, because, in the case of multiple alleles, the definition
of FST only holds true for the special case of random differentiation without selection (Nei,
1973). Both GST (e.g. Hamrick and Godt, 1997) and FST (e.g. Pressoir and Berthaud, 2004a)
have been used for studying crop genetic diversity. While these measures normally employ
14
molecular data, if additive allelic effects are assumed, an FST measure for quantitative
morphological traits can be derived using variance components (Pressoir and Berthaud, 2004a).
Alternatively, diversity may be usefully measured as genetic distance-similarity.
Mohammadi and Prasanna (2003) discuss commonly used measures of genetic similarity based
on molecular marker data (Nei and Li’s coefficient, Jaccard’s coefficient, simple matching
coefficient, and modified Roger’s distance) and report that modified Roger’s distance is
preferred due to its superior statistical and genetic properties.
Genetic Data
The above parameters require data at the genetic level. Techniques for obtaining such
data include morphological data, biochemical data, and molecular (genome-based) data
(Mohammadi and Prasanna, 2003). Pedigree data may be used as well but require the
assumption that parents of unknown parentage are genetically distinct (Witcombe, 1999).
Measuring morphological traits is the classical method but is indirect because trait
expression has both genetic and environmental components (Newbury and Ford-Lloyd, 1997).
Studies have used replicated agronomic trials to gather morphological data (e.g. Patra and Dhua,
1998; Louette et al., 1997), but it is also possible to use on-farm measurements for traits that are
minimally influenced by the environment (e.g. Salick et al., 1997).
Biochemical markers and molecular markers have become popular for estimating
diversity because they are relatively unaffected by the environment. Biochemical markers
include the use of protein markers such as isozymes (Brown, 1978) (different forms of an
enzyme coded for by a single locus) or chemicals such as terpenes (Glaubitz and Moran, 2000).
Biochemical markers have now been superceded by techniques that sample DNA directly
15
(Newbury and Ford-Lloyd, 1997). Several types of DNA-based markers are available including:
restriction fragment length polymorphisms (RFLPs), simple sequence repeats (SSRs), random
amplified polymorphic DNA (RAPDs), amplification fragment length polymorphisms (AFLPs),
and single nucleotide polymorphisms (SNPs). Glaubitz and Moran (2000) review many of these
techniques and their properties.
A number of studies have compared the use of different marker systems. In maize
populations RFLPs were found to detect more alleles per locus on average than isozymes,
although estimates of population differentiation based on this data were similar (Dubreuil and
Charcosset, 1998). Pejic et al. (1998) conducted a comparative study of the use of RFLPs,
RAPDs, SSRs, and AFLPs in maize inbred lines. They found that dendrograms of genetic
similarity were comparable for all techniques except RAPDs. SSRs provided the most
information about heterozygosity and allele number and AFLPs the least, however, AFLPs were
most efficient because they reveal several bands with a single assay (Pejic et al., 1998). Highthroughput methods involving bulking of individuals have been used for SSRs to increase their
efficiency (Warburton et al., 2001). The procedural components of these molecular techniques
are constantly being improved. Some examples include the development of fast extraction
techniques for DNA (Csaikl et al., 1998), high throughput procedures for developing
microsatellite (SSR) libraries (Connell et al., 1998), and cooler conductive media for rapid DNA
electrophoresis (Brody and Kern, 2004).
Sampling Approaches
Acquiring data requires sampling crop plants in the field. Zarin et al. (1999) outlined a
methodology for sampling species/varietal diversity using quadrats in farmers’ fields and
16
emphasized sampling field borders because they tend to be especially diverse. When using
quadrats for sampling many small quadrats will accurately estimate abundance for common
species, but it often results in incomplete species lists. Conversely, a few large plots result in
more complete species lists, but tend to overestimate the abundance of rare species and give
imprecise abundance estimates for common species (McCune and Lesica, 1992). Agricultural
systems are human managed and subject recall may be an alternative method for capturing
information (Ashby, 1990). Asking farmers what crop species they planted, amount of seed
sown, and spacing can provide information on species/varietal richness and evenness without
requiring time-consuming and laborious quadrat sampling.
Sampling genetic diversity in the face of high population-to-population differentiation
increases the value of prior empirical data (Hamrick and Godt, 1997). Marshall and Brown
(1975) mathematically developed sampling strategies for ex-situ collection missions based on the
conservation of all alleles occurring at greater than 5% frequency within a population, being
captured 95% of the time. Their calculations led to the conclusion that no more than 50
individual plants should be collected per population, and as many populations (represented by
sites) as possible should be sampled (Marshall and Brown, 1975). This approach has the aim of
capturing diverse alleles of use for breeding. If the aim is to assess diversity, fewer samples are
needed. CIMMYT evaluated maize diversity with SSRs using two bulks of 15 individuals
(Warburton et al., 2001). Thus, diversity estimates require the collection of fewer individuals.
Mating systems (self-pollinating, cross-pollinating) affect genetic diversity of crop
species, with cross-pollinating species generally being more diverse within than between
populations (Hamrick and Godt, 1997). However, in the case where self-pollinating crop
populations are made up of a number of distinct lines, diversity may also be greater within than
17
between populations (Brush, 2004; Bekele, 1983). Such observations inform sampling strategies
as to at what level sampling should be most intensive. Diversity estimates are expensive and
often planning is based on proxy measures (e.g. environmental diversity) (Marshall and Brown,
1975).
Analytical Statistics for Diversity Measures
After data have been collected and parameters have been calculated, analytical statistical
techniques can be used to compare diversity. Richness measures can be compared using
standard statistical tests such as t-tests, analysis of variance, generalized linear models, etc.
Comparison of diversity indices is slightly more complex because the distributional properties of
the index must be known before making comparisons (Hutcheson, 1970). Genetic parameters
can be compared in the same way as species richness using standard analytical statistical tests
(Hamrick and Godt, 1997).
Multivariate statistics, including clustering and ordination, are used to describe
both species diversity and genetic diversity. McCune and Grace (2002) reviewed and provided a
practical guide to the use of multivariate statistical techniques for analyzing species diversity.
Mohammadi and Prasanna (2003) provide a broad, well-referenced review of these techniques
for analyzing genetic diversity. Labate (2000) has described some of the software packages
available for analyzing diversity at the genetic level.
Functional Diversity
Not all diversity is useful. Functional diversity is the portion of diversity that is of use to
farmers or breeders. The level (alpha, beta, gamma) at which functional diversity is best
18
measured is debatable and likely context-dependent. Alpha diversity provides a good estimate of
how many varieties farmers use at the farm level. Beta diversity may be an appropriate diversity
measure where farmers use locally adapted varieties or where seed exchange results in rapid
varietal turnover. In a situation where a few individuals are responsible for maintaining and
redistributing diversity, as is the case of Amuesha shamans maintaining cassava diversity (Salick
et al., 1997), gamma diversity may be a pertinent measure of overall available diversity.
Genetic diversity is functional in terms of the traits it controls. The neutral theory of
molecular evolution proposed by Kimura (1968) hypothesizes that most genetic polymorphisms
(different alleles) have no adaptive significance. This suggests that most nucleotide substitutions
in a gene are functionally equivalent (Clegg, 1997). Thus, high levels of diversity at the
molecular level may not be a good indication of functional diversity. In maize, morphological
traits show much more differentiation between populations than molecular markers, because of
divergent selection of functional morphological traits, despite considerable gene flow among
populations (Pressoir and Berthaud, 2004a). Thus, morphological traits of biological importance
to farmers may be a superior measure of functional diversity than molecular markers.
The functionality of some traits may not be recognized unless appropriately challenged,
as in the case of disease resistance. Basing collection of diversity on only recognizable
morphological traits could miss dormant functional diversity. Molecular marker techniques
could be linked more directly to functional diversity by selecting markers that are known to lie
near or within genes that control specific traits of interest. Functional groups can be derived
from a matrix of traits of interest (Pillar, 1999). Conservation strategies could then be focused
on ensuring the conservation of functional groups. Unfortunately, functional groups will shift as
traits of interest change or as populations are added to the matrix of traits of interest.
19
Marshall and Brown (1975) divided alleles into four types based on being rare or
common and on being local or widespread. They argued that only local common alleles are
worth collecting because widespread alleles will be captured as a consequence and rare alleles
are usually deleterious and not worth collecting (Marshall and Brown, 1975). Charcosset and
Moreau (2004) suggested that conservation of functional diversity could be guided by using
molecular markers to reveal alleles that are not currently available in elite germplasm and thus
worth conserving. Focusing conservation on functional diversity, a constantly changing concept,
will involve subjective classification of what is considered functional.
COMPARISON OF IN-SITU AND EX-SITU CONSERVATION STRATEGIES
Advantages of In-Situ Conservation
Preservation of evolutionary processes (mutation, migration, recombination, selection) is
often cited as a major advantage of in-situ conservation (Brown, 1999). These processes
supposedly lead to the evolution of locally adapted crop varieties (Bellon et al., 1997), but
empirical evidence is lacking (Wood and Lenne, 1997). Local adaptation in composite crosses
has been observed to develop after 12 generations even without human selection (Suneson,
1956). Composite crosses may not be a good model for farmer varieties because compared to
farmer managed plant populations they have a much broader genetic base and survivorship is
simpler without selection pressure for non-environmental factors, such as quality, flavor, size,
etc. (Brown, 1999). Zeven (1996) reported that rapid and substantial changes occurred in a
wheat landrace grown outside of its native region in the 1920s. In-situ conservation can be a
backup to ex-situ collections in case of loss (Brush, 1999). Brush (1991) suggested that in-situ
20
conservation may be less expensive than ex-situ conservation, however this is not necessarily the
case (Smale et al., 2003).
Disadvantages of In-Situ Conservation
On-farm conservation of landraces by farmers, schools, and agricultural societies was
advocated in Europe in 1927, but these activities were unsuccessful due to World War II, lack of
funding, and variable teacher enthusiasm (Zeven, 1996). On-farm conservation was again
contemplated but dismissed during the conservation efforts of the 1970s and 80s. This was
because plant breeders wanted the conserved materials to be directly available to them and it was
also assumed that: 1) agricultural development would inevitably replace existing varieties with
modern ones, and 2) monetary compensation would be a necessary incentive for on-farm
conservation (Brush, 1999). It has also been assumed that farmers cannot be trusted to protect
important resources and that protecting these resources would condemn conservationist farmers
to perpetual poverty (Brush, 1991).
One constant criticism of in-situ conservation is that while it has been advocated, no
concrete framework has been developed for its implementation (Meng et al., 1998a; Brush,
1991). More recently, Maxted et al. (2002) and Bretting and Duvick (1997) outlined broad
methodologies for on-farm conservation projects. Projects must be flexible enough to adapt to
specific circumstances (Maxted et al., 2002) and this hinders the development of general
frameworks. In-situ projects have been initiated to test the implementation of in-situ
conservation and many of these projects have encountered logistical problems when working
with governments and farmers (Jarvis et al., 2000).
21
The perceived disadvantages may be overcome by a number of initiatives that make
conservation of agrobiodiversity more attractive through incentives. Bellon (2004) and Brush
(1999) outlined incentives and interventions that might facilitate on-farm conservation.
Interventions have been divided into market and non-market interventions (Brush, 1999) and/or
supply and demand interventions (Bellon, 2004).
Incentives for In-situ Conservation
Direct monetary incentives have been considered and implemented in developed
countries such as those of the European Union (FAO, 1998) and developing countries such as
Nepal, but it is unknown how long such subsidies can last (Bretting and Duvick, 1997).
Development of markets for the products of farmers’ varieties is a market or demand
intervention. Niche markets where farmers’ varieties are in demand can be identified and
constraints to market development (storage, transportation, information, marketing
incongruencies) can be alleviated (Brush, 1999). Meng et al. (1998a) found that the amount of a
traditional variety that was marketed was a strong predictor of its conservation. Developing a
market that previously did not exist is also possible, as demonstrated by the case of Cherokee
maize landraces sold as Indian maize flour (Brush, 1999). This case accessed the large lucrative
US market, a situation that is unlikely for crops in developing countries facing substantial market
barriers (Humpal and Guenette, 2000). Brush (2004) suggests that development of markets
along the line of specialty products (e.g. European appellation marketing) would require an
investment larger than the value of the crop diversity it would conserve.
The direct sale of genetic resources is also a potential market. Markets and institutions
have been developed to facilitate the sale of genetic resources to the pharmaceutical industry (de
22
Carvalho, 2003). While it is possible to imagine similar systems for agrobiodiversity, large
public collections and the propensity for breeders to work with their own materials makes
development of such a market unlikely (Brush, 1999). The University of California at Davis
attempted to compensate developing countries for crop germplasm, however, lack of clear
ownership makes it difficult to compensate individuals (Ronald, 1998).
Participatory crop improvement is often proposed as an incentive for farmers to conserve
crop diversity on farms (Bellon, 2004; Smale et al., 2003; Maxted et al., 2002; Brush, 1999) and
is a non-market or supply intervention. Two methods of participatory crop improvement have
been outlined: 1) participatory varietal selection (farmer evaluation of finished varieties) and 2)
participatory breeding (farmer selection within highly variable populations) (Morris and Bellon,
2004). Brush (2004) asserts that participatory varietal selection can reduce on-farm diversity by
replacing traditional varieties and suggests that emphasis should be on participatory breeding.
Regardless of the participatory breeding approach used, if use of locally adapted populations is
encouraged, so will diversity (Morris and Bellon, 2004).
Evidence from participatory rice variety development in Nepal suggests the products of
participatory breeding can be widely adopted (Joshi et al., 2001). This study noted competition
between released and traditional varieties but did not monitor changes in diversity. Participatory
breeding should aim to improve farmers’ livelihoods by providing access to appropriate
agricultural technology (acceptable varieties). Whether this leads to increased diversity will be
case-dependent.
Farmers experiment (Bellon, 2001; Richards, 1989; Lightfoot, 1987; Lightfoot, 1984;
Johnson, 1972). Farmer experimentation with varieties is expressed in high rates of varietal
turnover (Bellon et al., 1997). Increasing access to varietal diversity provides farmers with more
23
options for maintaining diversity during varietal turnover. Transaction costs (time, effort,
resources) for obtaining seed may be high (Bellon, 2004). Creation of an information-rich
environment in which farmers can obtain seed (e.g. CIMMYT’s maize landrace project in
Oaxaca, Mexico (Smale et al., 2003)), can facilitate the spread and use of crop genetic resources.
Seed exchange may also occur spontaneously at diversity fairs in which farmers gather at
central locations to display the crops and varieties they grow (Brush, 1999). Community
seedbanks have been established to facilitate access to varietal diversity (Cromwell and van
Oosterhout, 2000; Asfaw, 1999). Seed regulatory frameworks often require stringent levels of
uniformity and quality that prevent development of heterogenous materials (high diversity) and
restrict farmer participation (Wolff, 2004; Louwaars, 2001). Reform of these laws could
improve access. Education campaigns and publicity supporting the conservation of crop genetic
resources have also been used to promote conservation (Bretting and Duvick, 1997).
Advantages of Ex-Situ Conservation
The principal advantage of genebanks is that their materials are readily available to plant
breeders. Characterization and evaluation of materials and storage of this information in
databases also facilitate the process of plant breeding. These collections address the uncertainty
of what will be required in the future because they contain a broad range of materials (Smale and
Day-Rubenstein, 2002). Storage in genebanks guards against the loss of diversity in
agroecosystems (Zeven, 1996) and can facilitate reintroduction in the case of loss (FAO, 1998).
Reintroduction has become especially important in cases of disaster relief following war or
natural disasters (Wye-University, 2005). Ex-situ conservation also has the advantage of an
24
established theoretical basis that can guide decision-making regarding the collection,
characterization, and utilization of agrobiodiversity.
Disadvantages of Ex-Situ Conservation
Limited use of ex-situ collections by breeders has resulted in the questioning of
genebanks’ value (Wright, 1997). Plant breeders often prefer to work within their own materials
because they can still achieve genetic gains (Cooper et al., 2001) and crossing elite to
unimproved materials can degrade the genetic gains of elite breeding lines (Cuevas-Pérez et al.,
1992). In contrast, Smale and Day-Rubenstein (2002) have shown that germplasm requests from
US genebanks are substantial and developing countries are major recipients of distributed
materials. It has also been demonstrated that current and future benefits from stored germplasm
likely outweigh the costs of ex-situ conservation (Pardey et al., 1998).
Ex-situ conservation has been criticized as being static (Bellon et al., 1997; Brush, 1991).
Brush (1991) argues that if diversity is adequately collected today it will quickly become
obsolete due to evolution in agroecosystems. This is an extreme view because if a complete
collection were obtained formal breeding and introduction programs could mimic these in-situ
processes (e.g. crossing, migration, and selection). Even mutation can be induced (van Harten,
1998). These programs may not be as cost-effective at carrying out these processes as in-situ
conservation.
Collection missions of the IBPGR were thought to have preserved most of the crop plant
diversity available at the time (Plucknett et al., 1987). However, poor sampling techniques
(Frankel et al., 1995), incomplete data regarding agroecological setting (Bretting and Duvick,
1997), and the focus on a few major species (Lawrence, 1984) were associated with this effort.
25
Only as much as 50% of the genetic variation for minor crops has been sampled and passport
data are unavailable for 50% of all accessions (Wright, 1997).
Maintenance of collections is problematic. Regeneration of accessions is costly and
funding is often lacking. Regeneration backlogs in 66% of developing countries from 1995 to
2000 (Qualset and Shands, 2005). Worldwide only 18% of countries have been able to reduce
their regeneration backlogs (Wye-University, 2005). This has resulted in genebanks being
referred to as seed morgues (Goodman, 1990). Even if regeneration occurs, random genetic drift
is a concern in small populations and can affect allele frequency through loss of heterozygosity
and due to the fixation or loss of alleles (Yeh, 2000). Techniques such as those described by
Gale and Lawrence (1984) minimize these risks. These techniques do not address the problem of
initial samples being inadequately small. Also, farmers may not be able to provide seed from a
theoretically optimal number of individuals during collections (Mazzani and Segovia, 1998).
While the advantages seem few and the disadvantages seem many, ex-situ conservation
has been the mainstay of agrobiodiversity conservation. This is because it guarantees the
conservation of plant genetic resources and makes them available to breeding programs. Core
collections are reduced collections that contain most of the diversity present in all accessions.
They can be established on the basis of phenotypic or genetic variability present in all accessions
and are less costly to maintain because they consist of fewer accessions (Scippa et al., 2001).
Statistical techniques for the selection of core subsets have been developed (Franco et al., 2006;
Franco et al., 2005). Screening materials for traits of interest is costly. Optimum search
strategies have been outlined to facilitate the use of germplasm when variability that does not
exist within breeding lines is needed (Smale, 1998). Such strategies may improve the efficiency
of ex-situ conservation.
26
SEED EXCHANGE
Eighty percent of all seed in developing countries is produced on-farm and seed exchange
is one of the ways that local gene pools are maintained (Almekinders and de Boef, 1999).
Farming systems are dynamic and involve substantial amounts of seed exchange both within and
between communities (Louette et al., 1997). Seed exchange can occur across long distances
(Brush et al., 1981). Even small amounts of seed exchange can prevent genetic differentiation in
open-pollinated crops (Pressoir and Berthaud, 2004b). Still, adoption of seed declines as
distance from source increases (Witcombe et al., 1999), in part because of reduced seed
exchange. Local seed exchange is limited in its ability to acquire exotic materials (Almekinders
and de Boef, 1999). Exotic materials can increase diversity by being incorporated into a
farmer’s suite of varieties (Brush, 2004) or through creolization (repeated crossing of modern
variety to farmer varities) in the case of open-pollinated crops (Bellon and Risopoulos, 2001).
Seed exchange networks can facilitate recovery of lost varieties and a network of farmers
can maintain many varieties at a lower cost than can an individual farmer (Bellon, 2004). Seed
networks can be weak with regards to incentives, information, and resources (Tripp, 2001) but
can also be complex, dynamic and efficient (Cromwell, 1990). Certain farmers maintain larger
amounts of diversity than others (Meng et al., 1998a). Facilitating seed exchange between these
farmers and other farmers can enhance diversity (Cromwell and van Oosterhout, 2000). Farmer
seed exchange tends to be based on family ties and traditional social networks (Almekinders et
al., 1994). Thus, social barriers tend to prevent seed exchange (Zeven, 1999). Seboka and
Deressa (2000) argue that government extension programs should become involved in informal
27
seed exchange networks in order to validate them. Seed exchange is ubiquitous and is an
important mechanism underpinning defacto in-situ conservation.
ETHICS AND FARMERS’ RIGHTS
Intellectual property rights (IPRs) have become legally entrenched through the World
Trade Organization (WTO), which obligates members to accept its agreement on trade-related
aspects of intellectual property rights (TRIPs) and as such plant variety protection through
patenting (Alker and Heidhues, 2002). Brush (2004) traces the history of how genetic resource
status changed from that of common heritage to that of private property. Breeders’ rights have
been developing since the 1920s but were limited by breeders’ exemption (right to use varieties
in breeding) and farmers’ privilege (right to save and re-use seed) clauses, until 1991 when these
clauses were cut back (Wolff, 2004). Plant variety protection legally prohibits over-the-fence
exchange (i.e. from one farmer to another) of seed (Alker and Heidhues, 2002). However, it is
unclear if this is enforceable where informal seed networks is concerned. This is because many
users conducting relatively small transactions will likely not merit enforcement. Moreover,
developing countries have weak institutions for protecting such rights (Evenson, 1999).
A more realistic concern is that patented products originating from local genetic
resources and knowledge will not benefit local people, a phenomenon which has been termed
biopiracy (Shiva, 1997). Farmers’ rights have been viewed as a response to breeders’ rights
which permit proprietary claims to finished varieties (Brush, 1992). Farmers’ rights have also
been proposed as a means by which farmers in poor countries could be compensated for their
contribution to the development and maintenance of crop genetic resources (Esquinas-Alcázar,
1998). The ethical issues surrounding the use of plant genetic resources and adequate
28
compensation of those involved in their maintenance will continue to be an important issue in
crop genetic resources conservation.
CONCLUSIONS
Agrobiodiversity has become an international priority and is institutionalized through
binding international legal agreements. The value of agrobiodiversity is unquestionable, but will
also depend on the clientele (e.g. breeders, poor farmers, etc.). The measurement of
agrobiodiversity is a necessary prerequisite for developing conservation strategies. Measurement
can be carried out using a number of different methods for a number of different units of analysis
(species, varieties, genes). Use of specific methods, units of analysis, and sampling strategies
will depend on the objectives of the conservation program. Likewise, the statistical techniques
available for analysis of agrobiodiversity data are numerous and appropriate utilization is
dependent on objectives. Implementation of conservation strategies falls broadly into in-situ and
ex-situ approaches. While these approaches are deemed to be theoretically complementary, there
are few examples of projects implementing integrated approaches. Each of these strategies has
advantages and disadvantages, although some are currently debated. Despite these difficulties, it
is clear that the conservation of agrobiodiversity is a prerequisite for the development of
sustainable agricultural systems.
LITERATURE CITED
Alker, D., and F. Heidhues. 2002. Farmers' rights and intellectual property rights - reconciling
conflicting concepts, p. 61-92, In R. E. Evenson, et al., eds. Economic and social issues in
agricultural biotechnology. CAB International, London.
29
Almekinders, C., and W. de Boef. 1999. The challenge of collaboration in the management of
crop genetic diversity. ILEIA newsletter (December):5-7.
Almekinders, C., N.P. Louwaars, and G.H. de Bruijn. 1994. Local seed systems and their
importance for an improved seed supply in developing countries. Euphytica 78:207-216.
Altieri, M.A., and L.C. Merrick. 1987. In situ conservation of crop genetic resources through
maintenance of traditional farming systems. Economic Botany 41:86-96.
Asfaw, Z. 1999. The barleys of Ethiopia, p. 77-107, In S. B. Brush, ed. Genes in the field: onfarm conservation of crop diversity. International Plant Genetic Resources Institute
copublished with International Development Research Centre and Lewis Publishers,
Rome.
Ashby, J.A. 1990. Evaluating technology with farmers: a handbook. Centro International de
Agricultura Tropical, Cali.
Bateson, W. 1913. Mendel's principles of heredity University Press, Cambridge.
Bekele, E. 1983. Some measures of gene diversity analysis on land race populations of Ethiopian
barley. Hereditas 98:127-143.
Bellon, M.R. 1996. The dynamics of crop infraspecific diversity: a conceptual framework at the
farmer level. Economic Botany 50:26-39.
Bellon, M.R. 2001. Participatory research methods for technology evaluation: a manual for
scientists working with farmers CIMMYT, Mexico, D.F.
Bellon, M.R. 2004. Conceptualizing interventions to support on-farm genetic resource
conservation. World Development 32:159-172.
Bellon, M.R., and J. Risopoulos. 2001. Small-scale farmers expand the benefits of improved
maize germplasm: a case study from Chiapas, Mexico. World Development 29:799-811.
Bellon, M.R., J.-L. Pham, and M.T. Jackson. 1997. Genetic conservation: a role for rice farmers,
p. 261-289, In M. Maxted, et al., eds. Plant genetic conservation: the in situ approach.
Chapman & Hall.
Brennan, J.P., and D. Byerlee. 1991. The rate of crop varietal replacement on farms: measures
and empirical results for wheat. Plant Varieties and Seeds 4:99-106.
Bretting, P.K., and D.N. Duvick. 1997. Dynamic conservation of plant genetic resources.
Advances in Agronomy 61:1-51.
Brockway, L.H. 1979. Science and colonial expansion: the role of the British Royal Botanical
Gardens. Academic Press, New York.
Brody, J.R., and S.E. Kern. 2004. Sodium boric acid: a Tris-free, cooler conductive medium for
DNA electrophoresis. BioTechniques 36:214-216.
Brown, A.H.D. 1978. Isozymes, plant population genetic structure and genetic conservation.
Theoretical and Applied Genetics 52:145-157.
Brown, A.H.D. 1999. The genetic structure of crop landraces and the challenge to conserve them
in situ on farms, p. 29-48, In S. B. Brush, ed. Genes in the field: on-farm conservation of
crop diversity. Internaitonal Plant Genetic Resources Institute copublished with
International Development Research Centre and Lewis Publishers, Rome.
Brush, S.B. 1991. A farmer-based approach to conserving crop germplasm. Economic Botany
45:153-165.
Brush, S.B. 1992. Farmers' rights and genetic conservation in traditional farming systems. World
Development 20:1617-1630.
Brush, S.B. 1999. The issues of in situ conservation of crop genetic resources, p. 3-26, In S. B.
Brush, ed. Genes in the field: on-farm conservation of crop diversity. International Plant
30
Genetic Resources Institute copublished with International Development Research Centre
and Lewis Publishers, Rome.
Brush, S.B. 2004. Farmers' bounty: locating crop diversity in the contemporary world. Yale
University Press, New Haven.
Brush, S.B., and E. Meng. 1998. Farmers' valuation and conservation of crop genetic resources.
Genetic Resources and Crop Evolution 45:139-150.
Brush, S.B., H.J. Carney, and Z. Huaman. 1981. Dynamics of Andean potato agriculture.
Economic Botany 35:70-85.
Ceccarelli, S. 1996. Positive interpretation of genotype by environment interactions in relation to
sustainability and biodiversity, p. 467–486, In M. Cooper and G. L. Hammer, eds. Plant
adaptation and crop improvement. CAB International, Oxon, UK.
Ceccarelli, S., J. Valkoun, W. Erskine, S. Weigand, R. Miller, and J.A.G. van Leur. 1992. Plant
genetic resources and plant improvement as tools to develop sustainable agriculture.
Experimental Agriculture 28:89-98.
Charcosset, A., and L. Moreau. 2004. Use of molecular markers for the development of new
cultivars and the evaluation of genetic diversity. Euphytica 137:81-94.
Clegg, M.T. 1997. Plant genetic diversity and the struggle to measure selection. Journal of
Heredity 88:1-7.
Cleveland, D.A., D. Soleri, and S.E. Smith. 1994. Do folk varieties have a role in sustainable
agriculture? Bioscience 44:740-751.
Cleveland, D.A., D. Soleri, and E. Smith. 1999. Farmer plant breeding from a biological
perspective: implications for collaborative plant breeding CIMMYT, Mexico, D.F.
Condit, R., N. Pitman, E.G. Leigh, J. Chave, J. Terborgh, R.B. Foster, P. Núñez, S. Aguilar, R.
Valencia, G. Villa, H.C. Muller-Landau, E. Losos, and S.P. Hubbell. 2002. Beta-diversity
in tropical forest trees. Science 295:666-669.
Connell, J.P., S. Pammi, M.J. Iqbal, T. Huizinga, and A.S. Reddy. 1998. A high through-put
procedure for capturing microsatellites for complex plant genomes. Plant Molecular
Biology Reporter 16:341-349.
Cooper, H.D., C. Spillane, and T. Hodgkin. 2001. Broadening the genetic base of crop
production CABI Publishing, London.
Cromwell, E. 1990. Seed diffusion mechanisms in small farmer communities: lessons from Asia,
Africa and Latin America Overseas Development Institute, London.
Cromwell, E., and S. van Oosterhout. 2000. On-farm conservation of crop diversity: policy and
institutional lessons from Zimbabwe, p. 217-238, In M. Smale, ed. Farmers, gene banks
and crop breeding: economic analyses of diversity in wheat, maize, and rice. Kluwer
Academic Press, USA.
Csaikl, U.M., H. Bastian, R. Brettschneider, S. Gauch, A. Meir, M. Schauerte, F. Scholz, C.
Sperisen, B. Vornam, and B. Ziegenhagen. 1998. Comparative analysis of different DNA
extraction protocols: a fast, universal maxi-preparation of high quality plant DNA for
genetic evaluation and phylogenetic studies. Plant Molecular Biology Reporter 16:69-86.
Cuevas-Pérez, F.E., E.P. Guimaráes, L.E. Berrío, and D.I. González. 1992. Genetic base of
irrigated rice in Latin America and the Caribbean, 1971-1989. Crop Science 32:10541059.
Darwin, C. 1860. The origin of species by means of natural selection. Thomas Y. Crowell
Company, Publishers, New York.
31
Darwin, C. 1883. The variation of animals and plants under domestication. 2nd ed. D. Appleton
& Co., New York.
de Candolle, A. 1914. Origin of cultivated plants D. Appleton & Co., New York.
de Carvalho, A.P. 2003. Adding value to Brazilian biodiversity through biotechnology, p. 237243, In I. Serageldin and G. J. Persely, eds. Biotechnology and sustanable development:
voices of the south and north. CAB International, London.
Dickinson, A. 1967. Carl Linnaeus: pioneer of modern botany. F. Watts, New York.
Dubreuil, P., and A. Charcosset. 1998. Genetic diversity within and among maize populations: a
comparison between isozyme and nuclear RFLP loci. Theoretical and Applied Genetics
96:577-587.
Ellstrand, N.C. 2001. When transgenes wander, should we worry? Plant Physiology 125:15431545.
Esquinas-Alcázar, J. 1998. Farmers' rights, p. 207-217, In R. E. Evenson, et al., eds. Agricultural
values of plant genetic resources. CAB International, London.
Evenson, R.E. 1999. Intellectual property rights, access to plant germplasm, and crop production
scenarios in 2020. Crop Science 39:1630-1635.
Evenson, R.E., and D. Gollin. 1997. Genetic resources, international organizations, and
improvement in rice varieties. Economic Development and Cultural Change 34:471-500.
FAO. 1998. The state of the world's plant genetic resources for food and agriculture, p. 510.
Food and Agriculture Organization of the United Nations, Rome.
FAO. 2001. The international treaty on plant genetic resources for food and agriculture Food and
Agriculture Organization of the United Nations, Rome.
Fleuret, P., and A. Fleuret. 1980. Nutrition, consumption and agricultural change. Human
Organization 39:250-260.
Fowler, C., and T. Hodgkin. 2004. Plant genetic resources for food and agriculture: assessing
global availability. Annual Review of Environment and Resources 29:143-179.
Franco, J., J. Crossa, S. Taba, and H. Shands. 2005. A sampling strategy for conserving genetic
diversity when forming core subsets. Crop Science 45:1035-1044.
Franco, J., J. Crossa, M. Warburton, and S. Taba. 2006. Sampling strategies for conserving
maize diversity when forming core subsets using genetic markers. Crop Science 46:854864.
Frankel, O.H. 1954. The development and maintenance of superior genetic stocks. Heredity
4:89-102.
Frankel, O.H., A.H.D. Brown, and J.J. Burbon. 1995. The conservation of plant biodiversity
Cambridge University Press, Cambridge.
Gale, J.S., and M.J. Lawrence. 1984. The decay of variability, p. 77-101, In J. H. W. Holden and
J. T. Williams, eds. Crop genetic resources: conservation and evaluation. Allen and
Unwin, London.
Glaubitz, J.C., and G.F. Moran. 2000. Genetic tools: the use of biochemical and molecular
markers, p. 39-59, In A. Young, et al., eds. Forest conservation genetics: principles and
practice. CABI Publishing, New York.
Goodman, M.M. 1990. Genetic and germplasm stocks worth conserving. Journal of Heredity
81:11-16.
Hammer, K. 2003. A paradigm shift in the discipline of plant genetic resources. Genetic
resources and crop evolution 50:3-10.
32
Hamrick, J.L., and M.J.W. Godt. 1997. Allozyme diversity in cultivated crops. Crop Science
37:26-30.
Hanski, I., and D. Simberloff. 1997. The metapopulation approach, its history, conceptual
domain, and application to conservation Metapopulation biology. Academic Press, Inc.
Harlan, H.V., and M.L. Martini. 1936. Problems and results of barley plant breeding, p. 303-346,
In USDA, ed. USDA yearbook of agriculture. US Government Printing Office,
Washington, DC.
Harlan, J.R. 1972. Genetics of disaster. Journal of Environmental Quality 1:212-215.
Harlan, J.R. 1975. Our vanishing genetic resources. Science 188:618-621.
Hartl, D.L., and A.G. Clark. 1989. Principles of population genetics. 2nd ed. Sinauer Associates,
Inc., Sunderland.
Hawkes, J.G. 1977. The importance of wild germplasm in plant breeding. Euphytica 26:615-621.
Hubbell, S.P. 2001. The unified neutral theory of biodiversity and biogeography. Princeton
University Press, Princeton.
Humpal, D., and P. Guenette. 2000. Processed food safety in developing economies,
http://agecon.tamu.edu/iama/2000Congress/Forum%20%20Final%20PAPERS/Area%20IV/Guenette_Paul.PDF.
Hurlbert, S.H. 1971. The nonconcept of species diversity: a critique and alternative parameters.
Ecology 52:577-586.
Hutcheson, K. 1970. A test for comparing diversities based on the Shannon formula. Journal of
Theoretical Biology 29:151-154.
Iltis, H.H. 1974. Freezing the genetic landscape: the preservation of diversity in cultivated plants
as an urgent social responsibility of plant geneticists and plant taxonomists. Maize
Genetics Cooperation News Letter 48:199-200.
Jarvis, D., B. Sthapit, and L. Sears, (eds.) 2000. Conserving agricultural biodiversity in situ: a
scientific basis for sustainable agriculture, proceedings of a workshop, 5-12 July 1999,
Pokhara, Nepal. International Plant Genetic Resources Institute, Rome.
Johnson, A.W. 1972. Individuality and experimentation in traditional agriculture. Human
Ecology 1:149-159.
Joshi, K.D., B.R. Sthapit, and J.R. Witcombe. 2001. How narrowly adapted are the products of
decentralised breeding? The spread of rice varieties from a participatory plant breeding
programme in Nepal. Euphytica 122:589-597.
Just, R.E., and D. Zilberman. 1983. Stochastic structure, farm size, and technology adoption in
developing agriculture. Oxford Economic Papers 35:307-328.
Kimura, M. 1968. Evolutionary rate at the molecular level. Nature 217:624-626.
Labate, J.A. 2000. Software for population genetic analysis of molecular marker data. Crop
Science 40:1521-1528.
Lawrence, T. 1984. Collection of crop germplasm: the first 10 years, 1974-84 IBPGR
Secretariat, Rome.
Lewandowski, I., M. Hardtlein, and M. Kaltschmitt. 1999. Sustainable crop production:
definition and methodological approach for assessing and implementing sustainability.
Crop Science 39:184-193.
Lightfoot, C. 1984. On-farm experimentation in farming systems research. Farming Systems
Newsletter 17:12-16.
Lightfoot, C. 1987. Indigenous research and on-farm trials. Agricultural Administration and
Extension 24:79-89.
33
Louette, D. 1999. Traditional management of seed and genetic diversity what is a landrace?, p.
109-142, In S. B. Brush, ed. Genes in the field: on-farm conservation of crop diversity.
International Plant Genetic Resources Institute copublished with International
Development Research Centre and Lewis Publishers, Rome.
Louette, D., A. Charrier, and J. Bertaud. 1997. In-situ conservation of maize in Mexico: genetic
diversity and maize seed management in a traditional community. Economic Botany
51:20-38.
Louwaars, N.P. 2001. Regulatory aspects of breeding for diversity, p. 105-114, In H. D. Cooper,
et al., eds. Broadening the genetic base of crop production. CABI Publishing copublished
with Food and Agriculture Organization of the United Nations and International Plant
Genetic Resources Institute, UK.
Margalef, D.R. 1958. Information theory in ecology. Genetic Systematics 3:36-71.
Marshall, D.R., and A.H.D. Brown. 1973. Stability of performance of mixtures and multilines.
Euphytica 22:405-412.
Marshall, D.R., and A.H.D. Brown. 1975. Optimum sampling strategies in genetic conservation,
p. 53-80, In O. H. Frankel and J. G. Hawkes, eds. Crop genetic resources for today and
tomorrow: genetic variation in plant populations. Cambridge University Press,
Cambridge, U.K.
Matson, P.A., W.J. Parton, A.G. Power, and M.J. Swift. 1997. Agricultural intensification and
ecosystem properties. Science 277:504-509.
Maxted, M., B.V. Ford-Lloyd, and J.G. Hawkes. 1997. Complementary conservation strategies,
p. 15-40, In M. Maxted, et al., eds. Plant genetic conservation: the in-situ approach.
Chapman & Hall, London.
Maxted, M., L. Guarino, L. Myer, and E.A. Chiwona. 2002. Towards a methodology for on-farm
conservation of plant genetic resources. Genetic Resources and Crop Evolution 49:31-46.
Mazzani, E., and V. Segovia. 1998. Recolección de especies cultivadas en la ecorregión del Río
Atabapo del estado Amazonas, Venezuela. Plant Genetic Resources Newsletter 113:2226.
McCune, B., and P. Lesica. 1992. The trade-off between species capture and quantitative
accuracy in ecological inventory of lichens and bryophytes in forests in Montana.
Bryologist 95:296-304.
McCune, B., and J.B. Grace. 2002. Analysis of ecological communities. MjM Software Design,
USA.
McNaughton, S.J. 1977. Diversity and stability of ecological communities: a comment on the
role of empiricism in ecology. The American Naturalist 111:515-525.
Meilleur, B.A., and T. Hodgkin. 2004. In situ conservation of crop wild relatives: status and
trends. Biodiversity and Conservation 13:663-684.
Meng, E.C.H., J.E. Taylor, and S.B. Brush. 1998a. Implications for the conservation of wheat
landraces in Turkey from a household model of varietal choice, p. 127-142, In M. Smale,
ed. Farmers, gene banks and crop breeding: economic analyses of diversity in wheat,
maize, and rice. Kluwer Academic Press, USA.
Meng, E.C.H., M. Smale, M.R. Bellon, and D. Grimanelli. 1998b. Definition and measurement
of crop diversity for economic analysis, p. 19-31, In M. Smale, ed. Farmers, gene banks
and crop breeding: economic analyses of diversity in wheat, maize, and rice. Kluwer
Academic Press, USA.
34
Mohammadi, S.A., and B.M. Prasanna. 2003. Analysis of genetic diversity in crop plants salient statistical tools and considerations. Crop Science 43:1235-1248.
Morris, M.L., and M.R. Bellon. 2004. Participatory plant breeding research: opportunities and
challenges for the international crop improvement system. Euphytica 136:21-35.
Naeem, S., L.J. Thompson, S.P. Lawler, J.H. Lawton, and R.M. Woodfin. 1994. Declining
biodiversity can alter the performance of ecosystems. Nature 368:734-737.
Nei, M. 1973. Analysis of gene diversity in subdivided populations. PNAS 70:3321-3323.
Newbury, H.J., and B.V. Ford-Lloyd. 1997. Estimation of genetic diversity, p. 192-262, In M.
Maxted, et al., eds. Plant genetic conservation: the in-situ approach. Chapman & Hall,
London.
Pardey, P.G., B. Skovmand, S. Taba, M.E. van Dusen, and B.D. Wright. 1998. The cost of
conserving maize and wheat genetic resources ex-situ, p. 35-55, In M. Smale, ed.
Farmers, gene banks and crop breeding: economic analyses of diversity in wheat, maize,
and rice. Kluwer Academic Press, USA.
Patra, B.C., and S.R. Dhua. 1998. Collecting and evaluating rice germplasm for Orissa, India.
Plant Genetic Resources Newsletter 113:53-55.
Peet, R.K. 1974. The measurement of species diversity. Annual Review of Ecology and
Systematics 5:285-307.
Pejic, I., P. Ajmone-Marsan, M. Morgante, V. Kozumplick, P. Castiglioni, G. Taramino, and M.
Motto. 1998. Comparative analysis of genetic similarity among maize inbred lines
detected by RFLPs, RAPDs, SSRs, and AFLPs. Theoretical and Applied Genetics
97:1248-1255.
Pillar, V.D. 1999. On the identification of optimal plant functional types. Journal of Vegetation
Science 10:631-640.
Pistorius, R. 1997. Scientists, plants and politics: a history of the plant genetic resources
movement. International Plant Genetic Resources Institute, Rome.
Plucknett, D.L., N.J.H. Williams, and N.M. Anishetty. 1987. Gene banks and the world's food.
Princeton University Press, Princeton, N.J.
Pressoir, G., and J. Berthaud. 2004a. Population structure and strong divergent selection shape
phenotypic diversification in maize landraces. Heredity 92:95-101.
Pressoir, G., and J. Berthaud. 2004b. Patterns of population structure in maize landraces from the
Central Valleys of Oaxaca in Mexico. Heredity 92:88-94.
Purvis, A., and A. Hector. 2000. Getting the measure of biodiversity. Nature 405:212-219.
Qualset, C.O., and H.L. Shands. 2005. Safeguarding the future of U.S. agriculture: the need to
conserve threatened collections of crop diversity worldwide University of California,
Division of Agriculture and Natural Resources, Genetic Resources Conservation
Program, Davis, CA, USA.
Qualset, C.O., A.B. Damania, A.C.A. Zanatta, and S.B. Brush. 1997. Locally based crop plant
conservation, p. 160-175, In M. Maxted, et al., eds. Plant genetic conservation: the in-situ
approach. Chapman & Hall, London.
Richards, P. 1989. Farmers also experiment: a neglected intellectual resource in African science.
Discovery and Innovation 1:19-25.
Ronald, P.C. 1998. The genetic resources recognition fund. AgBiotech News and Information
10:19N-22N.
Ryerson, K.A. 1933. History and significance of the foreign plant introduction work of the
United States department of agriculture. Agricultural History 7:110-128.
35
Salick, J., N. Cellinese, and S. Knapp. 1997. Indigenous diversity of cassava: generation,
maintenance, use and loss among the Amuesha, Peruvian Upper Amazon. Economic
Botany 51:6-19.
Scippa, G., G.B. Polignano, and P. Uggenti. 2001. Diversity analysis and core collection
formation in Bari faba bean germplasm. Plant Genetic Resources Newsletter 125:33-38.
Seboka, B., and A. Deressa. 2000. Validating farmers' indigenous social networks for local seed
supply in Central Rift Valley of Ethiopia. Journal of Agricultural Education and
Extension 6:245-254.
Shiva, V. 1997. Biopiracy: the plunder of nature and knowledge. South End Press, Boston.
Smale, M. 1997. The green revolution and wheat genetic diversity: some unfounded
assumptions. World Development 25:1257-1269.
Smale, M., (ed.) 1998. Farmers, gene banks and crop breeding: economic analyses of diversity in
wheat, maize, and rice. Kluwer Academic Publishers, USA.
Smale, M., and K. Day-Rubenstein. 2002. The demand for crop genetic resources: international
use of the US national plant germplasm system. World Development 30:1639-1655.
Smale, M., M.R. Bellon, J.A. Aguirre, I. Manuel-Rosas, J. Mendoza, A.M. Solano, R. Martinez,
A. Ramirez, and J. Berthaud. 2003. The economic costs and benefits of a participatory
project to conserve maize landraces on farms in Oaxaca, Mexico. Agricultural Economics
29:265-275.
Suneson, C.A. 1956. An evolutionary plant breeding method. Agronomy Journal 48:188-191.
Swanson, T., and T. Goeschl. 2000. Optimal genetic resource conservation: in-situ and ex-situ, p.
165-191, In S. B. Brush, ed. Genes in the field: on-farm conservation of crop diversity.
IPGRI, Rome.
ten Kate, K., and S.A. Laird. 2000. Introduction, p. 1-12, In K. ten Kate and S. A. Laird, eds. The
commercial use of biodiversity: access to genetic resources and benefit-sharing.
Earthscan Publications Ltd., London.
Tilman, D. 1998. The greening of the green revolution. Nature 396:211-212.
Tilman, D., D. Wedin, and J. Knops. 1996. Productivity and sustainability influenced by
biodiversity in grassland ecosystems. Nature 379:718-720.
Trenbath, B.R. 1974. Biomass productivity of mixtures. Advances in Agronomy 26:177-210.
Tripp, R. 2001. Seed provision and agricultural development. Overseas Development Institute,
London.
UNCED. 1992. Convention on biological diversity United Nations Conference on Environment
and Development, Geneva.
van Harten, A.M. 1998. Mutation breeding: theory and practical application. Cambridge
University Press, UK.
Vavilov, N.I. 1951. The origin, variation, immunity, and breeding of cultivated plants. Chronica
Botanica 13:1-366.
Vavilov, N.I. 1957. Agroecological survey of the main field crops. The Academy of Sciences of
the USSR, Moscow.
Vellend, M. 2001. Do commonly used indices of ß-diversity measure species turnover? Journal
of Vegetation Science 12:545-552.
Warburton, M., X. Xianchun, S. Ambriz, L. Diaz, E. Villordo, and D. Hoisington. 2001. Use of
molecular markers in maize diversity studies at CIMMYT, pp. 130-133 Seventh Eastern
and Southern Africa regional maize conference, 11th - 15th February.
36
Weir, B.S., and C.C. Cockerham. 1984. Estimating F-statistics for the analysis of population
structure. Evolution 38:1358-1370.
Whittaker, R.H. 1965. Dominance and diversity in land plant communities. Science 147:250260.
Wilkes, H.G. 1993. Gerplasm collections: their use, potential, social responsibility, and genetic
vulnerability, p. 445-450, In D. R. Buxton, et al., eds. International crop science I. Crop
Science Society of America, Madison, WI.
Wilson, E.O., and P.R. Ehrlich. 1991. Biodiversity studies: science and policy. Science 253:758762.
Witcombe, J.R. 1999. Does plant breeding lead to a loss of genetic diversity?, p. 245-272, In D.
Wood and J. M. Lenne, eds. Agrobiodiversity: characterization, utilization, and
management. CABI Publishing, New York.
Witcombe, J.R., R. Petre, S. Jones, and A. Joshi. 1999. Farmer participatory crop improvement.
IV. The spread and impact of Kalinga III - a rice variety identified by participatory
varietal selection. Experimental Agriculture 35:471-487.
Wolff, F. 2004. Legal factors driving agrobiodiversity loss. Environmental Law Network
International 1:1-11.
Wood, D., and J.M. Lenne. 1997. The conservation of agrobiodiversity on-farm: questioning the
emerging paradeigm. Biodiversity and Conservation 6:109-129.
Wood, D., and J.M. Lenne. 1999. Agrobiodiversity: characterization, utilization, and
management CABI Publishing, New York.
Wright, B.D. 1997. Crop genetic resource policy: the role of ex situ genebanks. The Australian
Journal of Agricultural and Resource Economics 41:81-115.
Wright, S. 1951. The genetical structure of populations. Annals of Eugenics 15:323-354.
Wye-University, (ed.) 2005. Crop diversity at risk: the case for sustaining crop collections, pp. 133. Imperial College Wye, UK.
Yang, R.-C., and C. Yeh. 1992. Genetic consequences of in situ and ex situ conservation of
forest trees. The Forestry Chronicle 68:720-729.
Yeh, F. 2000. Population genetics, p. 21-37, In A. Young, et al., eds. Forest conservation
genetics: principles and practice. CABI Publishing, New York.
Zarin, D.J., G. Huijin, and L. Enu-Kwesi. 1999. Methods for the assessment of plant species
diversity in complex agricultural landscapes: guidelines for data collection and anlysis
from the PLEC Biodiversity Advisory Group (BAG). PLEC News and Views 13:3-16.
Zeven, A.C. 1996. Results of activities to maintain landraces and other material in some
European countries in situ before 1945 and what we may learn from them. Genetic
Resources and Crop Evolution 43:337-341.
Zeven, A.C. 1998. Landraces: a review of definitions and classifications. Euphytica 104:127139.
Zeven, A.C. 1999. The traditional inexplicable replacement of seed and seed ware of landraces
and cultivars: a review. Euphytica 110:181-191.
Zimmerer, K.S. 1991. Labor shortages and crop diversity in the Sourthern Peruvian Sierra.
Geographical Review 81:414-432.
37
Download